CMOS-Compatible Fluidic Chip Cooling Using Buried Channel Technology

advertisement
CMOS-Compatible Fluidic Chip Cooling Using
Buried Channel Technology
by
Daniel M. Gaugel
A thesis submitted in partial fulfillment of the requirements for
the degree of
Master of Science in
Electrical & Computer Engineering
Carnegie Mellon University
Pittsburgh, Pennsylvania
USA
Advisor: Professor Kaigham J. Gabriel
Acknowledgements
The author would like to thank the entire MEMS Laboratory at Carnegie Mellon
University, Ken Gabriel, Mary Moore, Drew Danielson, Elaine Lawrence, Lynn Philibin,
the staff of the CMU Nanofabrication Facility (Chris Bowman, Carsen Kline, and Tim
Fisher), the staff of the ECE Machine Shop (Jim Schubert, Dave Belotti, and Steve
Hoffman), Hasnain Lakdawala, Xu Zhu, George Lopez, John Neumann, Tamal
Mukherjee, Keith Rebello, Kyle Lebouitz, Suresh Santhanam, Joe Suhan, Jeff Zahn, Lars
Erdmann, and Chi-Fu Wu for their assistance and advice throughout the duration of this
project.
Table of Contents
I.
Introduction
1
II.
Background
2
III.
Fabrication
5
IV.
Thermal Analysis
i.
Simulation Model
14
ii.
Experimental Setup
20
iii.
Modeled, Experimental Results
21
V.
CMOS Compatibility
27
VI.
Conclusions
29
VII.
References
32
VIII.
Appendix A (Processing recipe)
33
IX.
Appendix B (Data acquisition system photographs)
34
I. Introduction
Electronic system power dissipation has increased exponentially since 1980, driven
primarily by the increase in transistor switching frequency. While power per gate continues to
fall (0.64-0.15 µW/ MHz from 1995-99) due to voltage reduction, overall power dissipation has
increased in virtually every electronic product due to increasing transistor density and frequency
[13]. For example, Intel microprocessor thermal design power has risen from 40.0 W for the
Pentium II [14] to 75.3 W for the current Pentium 4 [15]. Should this trend continue to rise,
forced air convection will no longer offer adequate electronic thermal management and Moore’s
Law will come to an end. Microfluidic convective heat transfer offers an alternative to the
conventional heat sink air-cooling systems. The superior fluid properties of water compared to
air (approximately 4 and 1,000 times greater specific heat and density, respectively [9]) offer
more efficient cooling within a compact volume. Designs featuring fluidic heat transfer
integrated into the silicon substrate offer high cooling capacity due to the high heat transfer
characteristics of micron-order diameter cooling channels. Significant fabrication challenges are
present with this approach however, including the need for system level and CMOS processing
integration.
We present a CMOS-compatible method of fabricating buried channels within bulk
single-crystal silicon serving as the substrate of CMOS circuitry. The channels were designed to
provide a compact, forced convection heat transfer liquid cooling approach to microprocessor and
integrated circuit thermal management. The baseline fabrication process consumes less than 8 %
of surface area and serves as a foundation for further area reduction, potentially leading to
incorporation of buried-channel cooling with CMOS electronics. A microchannel heat transfer
solution was designed, simulated, processed, and tested with a liquid cooling data acquisition
system. Compatibility with CMOS was characterized by evaluating the post-process transistor
performance of a simple oscillator circuit.
II. Background
The goal of integrated forced convection fluidic microprocessor cooling has provided a
fertile area of research into microchannel development.
D.B. Tuckerman [1] achieved heat removal between 181-790 W/cm2 with high aspect
ratio rectangular channels in 1981. Internal heat transfer enhancement through the reduction of
thermal resistance between fluid and substrate serves as the primary objective of this research. 50
µm wide by 300 µm deep vertical channels were processed within a silicon die with an
anisotropic KOH (potassium hydroxide) wet etch. The channels were spaced 50 µm apart over
the 1 cm2 surface area. The rectangular channel height/width aspect ratio ranged from 5-6 and
offered increased surface area with relatively small hydraulic diameters. A cover plate anodically
bonded to the fin array contained the fluid, with heat provided via thin film WSi2 resistors on the
chip surface. Experiments conducted with water at flow rates from 4.7-8.6 mL/min yielded fluid
pressures from 15-31 psig. The major disadvantage to this approach as a microprocessor cooling
solution lies in the wafer bonding, which will add processing steps and decrease yield.
Significant work on the development of optimum heat exchanger design has been
reported. By taking advantage of the high heat transfer coefficients of sub-cooled boiling,
Bergles et al [2] explored the design parameters to achieve optimum heat exchanger design as
well as summarized seminal two-phase heat exchange microchannel research. The authors
present a number of design recommendations for surface heat source, rectangular channel bulk
silicon substrate cooling die formations. These include locating the onset of incipient boiling
(requiring microscopic viewing of exiting fluid), establishing uniform flow distribution through
the channels, including header orifices to reduce the effect of pressure drop due to channel wall
friction, and spacing channel location for uniform substrate heat distribution.
de Boer [3] et al have reported buried channel fabrication using a single mask process
without the need for wafer bonding, eliminating alignment error. The buried channel technology
(BCT) involves a multi-step process in which access trenches are deep etched into the silicon
substrate. Microchannels are formed by isotropically etching the exposed silicon on the bottom
of the trench with the use of masked sidewalls. Two sidewall-masking methods (thermal SiO2
and LPCVD SixNy) successfully protected the trench sidewalls from the remaining etching steps.
An anisotropic etch cleared the SiO2 and SixNy at the base of the trenches, allowing channel
formation from isotropic silicon etching. A number of sidewall protection geometries are
presented, as well as a method of sealing and releasing the channels (Figure 1). Due to the high
temperature oxide growth, integration of CMOS electronics must be added after the formation of
the channels.
Joo et al [4] present a single-sided, low temperature, single mask microchannel
fabrication process. Rectangular cavities are formed by spinning a 10-µm layer of photoresist
onto a seed layer of nickel and chromium. After development, nickel-electroplating formed
channel walls in the areas not covered with the photoresist pattern. A subsequent electroplating
step after stripping the photoresist sealed the openings, leaving an enclosed microchannel on top
of the silicon substrate (Figure 2). The ability to grow the channels on top of electronics, with
thermal resistance considerably smaller than bonded heat sinks, present the major advantage of
this approach. The overall dimensions of the formed channels were 10 µm in width and 8 µm
high, with wall thickness of 10 µm.
III. Fabrication
The primary goal of this research effort lies in the development of a CMOS compatible
microchannel baseline fabrication process with the following features: 1) minimal mask area to
facilitate integration with electronics 2) channels located sufficiently deep into the substrate to
preserve electronics functionality 3) low temperature sidewall passivation (non-SiO2 growth).
Because polymers are resistant to etching inXeF2, we set out to create deep channels
using the passivation growth of C4F8. Anisotropic trenches could be etched using the Bosch
etch/passivation ICP (inductively coupled plasma) process, followed by an extended polymer
deposition. A directional removal of the polymer would expose the bulk silicon at the bottom of
the trench, allowing a spherical, bulbous pattern to be produced at the base using isotropic silicon
XeF2 etching. A series of these features patterned optimally close together would produce
connected spherical cavities resulting in a scalloped-wall buried sub-circuit channel. Due to the
size of the “access” mask area relative to the channel diameter, we could seal the access ports
using an extended polymer deposition. Figure 3 presents a conceptual drawing of the target
microchannel process.
While the eventual application for the microchannel process lies in integration with CMOS
electronics, silicon wafer samples were used to establish the baseline process without the need for
expensive and long-lead time microelectronics fabrication. A series of 4” masks were created
with various access-hole feature sizes and configurations to determine the final fabrication
baseline process.
We developed an eleven (11) step process to create buried microchannels in bulk silicon.
The procedure is presented schematically in Figure 4, with Appendix A listing the process steps
in greater detail, including plasma pressures and flow rates.
Figure 4: Processing recipe schematic
The processing steps presented in Figure 4 are as follows:
Steps 1-2) Photolithography, mask selection
Channel access patterns were formed on 1.5 cm x 1.5 cm SiO2-coated silicon chips with a
single spin, exposure, and development procedure. AZ 4620 photoresist was selected due to the
relatively thick film (~13 µm) required to maintain the mask pattern. This thickness restricted the
minimum feature sizes to just below 10 µm, but these feature sizes were later determined to be
sufficient for channel creation.
A sufficiently stable mask was needed to maintain the structural integrity of the pattern.
The low etch selectivity of photoresist to silicon (50-75/1) led to the exploration of stronger films
capable of withstanding the subsequent processing steps. We attempted to take advantage of the
high aluminum/silicon etch selectivity by patterning a single layer CMOS mask (0.44 µm Al +
1.25 µm SiO2). This was eliminated when excessive aluminum-fluoride reaction led to polymer
redeposition and micromasking during standard silicon etching (cross section presented in Figure
5). We selected a 4 µm SiO2 etch mask due to the successful results achieved with 1.25 µm SiO2
samples.
Step 3) Dry RIE SiO2 etch
A dry SiO2 etch was then used to develop the channel mask. Non-uniform SiO2 thickness
and over-etching caused the mask features to expand during this step, but recipe optimization
controlled the growth to 1-2 µm at most.
Step 4) Anisotropic etch
Next, the access trenches were created using the Bosch DRIE (deep reactive ion etch)
process. The trench depth dictated the maximum channel diameter, due to the isotropic
expansion of the channel. The trenches were etched 110-150 µm deep to leave a minimum of 50
µm of silicon between the top edge of the channel and the die surface.
Step 5) Isotropic RIE etch
An isotropic etch undercut the walls of the trenches 4-5 µm underneath the SiO2 mask
(cleaved cross section presented in Figure 6). We added this step to protect the upper portion of
the trenches during subsequent processing. The upper sections of the trenches are vulnerable in
the final channel etch (Step 9) due to the weakening of the sidewall protection during passivation
removal (Step 7). We hypothesized that the polymer along the top edges of the trenches
weakened due to the sidewalls lying inline to the path of the plasma ions. The plasma would
remove the polymer at the base of the trench, but weaken the protection along the upper sidewalls
in the process. During the XeF2 etch, the weakened polymer mask experienced undesirable
silicon sidewall etching (Figure 7), which diluted the process gases and increased the sacrificial
mask area.
Two methods were tested to alleviate the upper trench degradation. We attempted an
undercut demonstrated by de Boer [3] prior to the DRIE trench etch (Figure 8). After passivation,
the undercut area would thus be protected from the plasma by the mask. We witnessed sidewall
etching results similar to Figure 7. The second method, an isotropic etch between DRIE and
polymer deposition, was the most successful. The undercut mask blocked the subsequent plasma
ions from the upper trench area, protecting the polymer layer.
Steps 6-7) Polymer deposition, bottom trench polymer removal
Next, an 8.5-long minute deposition coated the trenches with a ~0.8-1.0 µm of C4F8
polymer. We then explored four methods to remove the polymer at the base of the trenches. The
first attempt involved an etch cycle of SF6 and C4F8 similar to the Bosch process, with
etch/passivation steps of 20-30 seconds. The second method involved a similar cycle,
substituting O2 for SF6 [18]. Both procedures failed to build a polymer sidewall mask that would
survive the remaining steps of the recipe. This is most likely due to the isotropic degradation of
the polymer with SF6 and O2. We thus attempted a similar etch/passivation cycle with Ar and
C4F8. While the Ar did not react with the polymer, the low power (30 W) plasma produced
scattering of ions, which weakened the polymer buildup along the sidewalls (Figure 9). Stronger
argon plasmas were therefore needed to produce anisotropic ion milling. Due to the control
restrictions our DRIE machine (STS) recipe, high power argon plasma could not be produced
using the etch/passivate cycle method. We found the most successful procedure for this step to
be an extended passivation followed immediately by an extended high power Ar plasma [19].
Increasing the platen power in the STS, which increased the impinging ion energy into the
polymer film, optimized the process. We found additional improvements by lowering the plasma
pressure, which increased the mean free path of the argon ions. A ramping procedure was needed
to build the argon plasma from 30 W platen power, 7 mT pressure at the onset to the optimal 150
W and ~1.5 mT conditions.
The Ar passivation removal step was difficult to control and reproducibility was poor.
This is most likely attributed to polymer deposition rate variations and curing. For an extended
8.5 minute-deposition, measured polymer thickness varied up to 20%. The polymer also cured
under vacuum, causing a range of argon milling times to effectively remove the polymer at the
base of the trench. Due to curing, we were unable to sufficiently clean the chamber before the
remaining STS recipe steps were completed. We observed that polymer removal at the base of
the trenches was best determined by aborting the process at the point where the bias voltage of
the platen would rapidly increase. We successfully estimated this point by recording the bias
voltage of a photoresist film silicon wafer and comparing to the ion milling process conditions.
Step 8) RIE etch
A 20 second-SF6 etch ensured the exposure of silicon at the base of the trench [17]. Nonuniformity observed in the channel diameters led to the inclusion of this step. Without this step,
sputtered polymer particles present at the base of the etch trench caused significant masking
during XeF2 etching.
Step 9) XeF2 etch
After a 30 minute-200 oC bake to cure the sidewall polymer [20], the channels were
formed at the base of the trench with an isotropic XeF2 etching process. Channel diameter was
controlled by adjusting the etch duration. Polymer delamination was evident at the base of the
trench walls, but the upper wall integrity remained intact. Figures 10-15 present cross sectional
and surface views of three samples processed for this analysis.
Steps 10-11) Oxygen polymer clearing, sealing
Oxygen plasma cleared the remaining polymer from the walls of the trenches.
At this point, there were two options for sealing the microchannels. An extended
polymer deposition could seal the access trenches at the mask surface if the access trenches were
reasonably small. Due to the 22-25 µm mask widths of the samples, a wafer bonding process was
implemented. We sealed the samples by bonding silicon sections of similar size to the openings
of the channels with a high temperature, water resistant epoxy.
With the process recipe yielding repeatable results, the final challenge lay in the
determination of the smallest successful feature size. We established the process with 300 µm
open areas and adjusted parameters for smaller trench areas. A 20 µm-wide trench proved to be
the smallest access feature size achieved during this work. Attempts to process channels through
smaller access holes were unsuccessful due a number of possible reasons, including insufficient
deposition growth and an inability to clear the polymer at the base of the holes.
An additional trend occurred during the XeF2 etch step that may qualify the worthiness of
the C4F8 polymer as an etch mask. Delamination was apparent on every test sample, with an
undesirable undercut occurring at the base of sidewalls. Figure 16 presents an open area
delamination indicative of this occurrence. The polymer may be permeable at weakened or thin
mask areas, which present the need for further processing analysis to ensure uniform, thick
sidewalls on the access trenches.
IV. Thermal Analysis
i. Simulation Model
The analysis of the heat sink capability of the processed buried channel samples presents
the second research objective. Performance models based on fundamental heat transfer principles
were created to examine the effects of channel geometry, flow rate, and chip substrate
temperature on heat transfer performance. While significant cooling advantages exist by
operating the system in dual phase (heat transfer due to change from liquid-gaseous state under
constant temperature), this analysis will focus on single-phase, sensible (change in coolant
temperature) heating of the coolant. The following analysis is derived from Chapters 3 and 8 of
[9].
An energy balance between the heat source (microprocessor), and the heat sink (fluid
passing through the channels) determines how the mean coolant temperature varies with position
along the channel. This balance also relates the heat transfer rate to the coolant inlet and outlet
temperatures. For the entire channel, irrespective of fluid transport properties or surface
conditions, this balance is expressed as:
Q"chip Achip = mcp∆Tcoolant (W)
with Q"chip = thermal heat load density (W/cm2), Achip = chip area (cm2), m = coolant mass flow
(kg/s), cp = coolant specific heat (J/kg K), and ∆Tcoolant = (Tcoolant, exit – Tcoolant, entrance).
It is necessary to relate the thermal heat density produced by the chip to the heat flux
from the substrate into the individual channels. The chip surface area relative to the cooling
channel surface area is therefore presented:
Achip = LengthchipWidthchip (cm2)
Achannels = (#channels)PerimeterchannelLengthchannel (cm2)
Q"chip Achip = Q"channels Achannels
The quantification of the forced convection between the chip substrate and the coolant is
expressed through Newton’s Law of Cooling for internal flow:
Q"channels = h(Tsubstrate − Tm ,coolant ) (W/cm2)
With Tsubstrate = substrate temperature (K), Tm,coolant = mean coolant temperature (K), and h = local
internal heat transfer coefficient (W/ m2 K), a parameter which encompasses all of the convection
heat transfer factors. These include boundary layer conditions (reliant on channel geometry), and
fluid thermal and transport properties. The convection heat flux is termed as positive when heat
is transferred from the surface to the fluid. With heat transfer occurring across the channel, the
mean coolant temperature will increase in the direction of the moving fluid.
Newton’s Law of Cooling applies to the case of a constant surface heat flux, with the
surface and fluid temperatures changing along the path of the channel. A variation exists for the
constant surface temperature, variable heat flux scenario:
q channels = h Achannels ∆Tlm (W)
with h = average internal heat transfer coefficient and ∆Tlm = log mean temperature difference,
which expresses the average temperature difference along the length of the channel. This is
defined as:
∆Tlm =
∆To − ∆Ti
(K)
ln(∆To / ∆Ti )
with ∆To = Tsurface – Tmean coolant outlet and ∆Ti = Tsurface – T mean coolant inlet. While a constant heat flux
would apply for the microprocessor model, the constant temperature model will be used for the
following analysis. The heat sink/ temperature controller used to simulate transistor thermal heat
generation will produce experimental results more accurately represented by the constant surface
temperature model. Figure 17 presents a simulated temperature profile along the length of the
channel for each case.
Figure 17: Axial temperature variations for heat transfer in a channel: a) constant surface heat flux,
b) constant surface temperature. Derived from [9].
In order to estimate the internal flow heat transfer coefficient and pumping power, we
analyzed the thermal and transport properties of the fluid. For the transport properties, the
Reynolds # represents the ratio of the inertial to viscous forces in the velocity boundary layer.
For internal flow through a circular channel, this parameter is expressed as:
Re# =
ρVL
µ
with µ = fluid absolute viscosity (N s/m2), ρ = fluid density (kg/m3), V = mean fluid velocity
(m/s), L = channel length (m). For laminar, or highly ordered and streamlined flow, Re # <
2,300.
The Nusselt # provides a measure of the convection heat transfer occurring at the surface.
This is expressed as a dimensionless temperature gradient:
∂
Nu # =
(T − Tsurface )
(T∞ − Tsurface )
y
∂
L
with T = fluid temperature within the boundary layer, and T∞ = fluid constant temperature outside
of the thermal boundary layer. For constant surface temperature and laminar, fully developed
conditions, the Nusselt # is constant and independent of channel location:
Nu # =
hD
= 3.66
k fluid
with D = channel diameter (m) and kfluid = fluid thermal conductivity (W/ m K). The heat transfer
coefficient and channel diameter are therefore inversely proportional, with greater heat transfer
occurring at smaller diameters.
Due to the non-uniform profiles of the buried channels, a hydraulic, or effective, radius is
needed to calculate the hydrodynamic and thermodynamic properties of the coolant. This is
defined as the ratio between the perimeter and the area of the channel. Effective diameter would
therefore be expressed as:
Deff = 4
Areachannel
(m)
Perimeterchannel
An iterative analysis can now be performed by equating the energy balance with the
constant temperature internal flow model:
qconv = h Achannels ∆Tlm = mcp(Tcoolant ,out − Tcoolant ,in ) (W) (1)
The thermal heat load can therefore be estimated by specifying a range of surface
temperatures, channel diameters, and flow rates, and iterating each case with the coolant outlet
temperature.
These relations are strictly confined for the fully developed regions of the internal
boundary layer region. The fully developed region (Figure 18) is defined as the point in the
channel where the velocity profile of the boundary layer no longer varies with increase in channel
distance. This point, referred to as the hydrodynamic entry length, can be estimated from the
Reynolds # and channel diameter for the laminar case:
x fd , h
D
≈ 0.05 Re
la min ar
Figure 18: Laminar, hydrodynamic boundary layer development in a circular channel. Derived
from [9].
A thermally fully developed condition is likewise represented as the point in the channel
where the fluid temperatures along the thermal boundary layer no longer change with channel
length. The thermal entry length can be estimated as:
x fd ,t
D
≈ 0.05 Re Pr
la min ar
with Pr = Prandtl #, defined as a ratio of momentum and thermal diffusivities. With water at 2290 oC, the Prandtl # varies from ~7-2, respectively. Therefore, hydrodynamic fully developed
conditions will arise closer to the channel entry than the thermal fully developed case. Figure 19
displays the cross section of the thermal boundary layer.
Figure 19: Thermal boundary layer development in a heated circular tube. Derived from [9].
The coolant flow pumping power presents an additional design variable for the
microchannel heat exchanger. Fluidic pressure differential across the chip is primarily attributed
to friction loss through the channel and entrance losses from the transition of fluid from the
header to the individual pathways. Channel pressure loss is expressed as:
∆P =
f ρV 2 L
(Pa)
2 Deff
With f = Moody (or Darcy) friction factor, a dimensionless measure of pressure drop for internal
flow. Similar to a voltage potential across parallel branches of an electrical circuit, the pressure
drop across the parallel channels will be identical for even flow distribution. The Moody friction
factor is a function of the surface roughness relative to the channel diameter and the Reynolds #:
f =
− (dp / dx)2 D
ε
= F (Re, )
2
d
ρV
with dp/dx = pressure gradient along the channel wall, and ε = wall roughness height (m). For
laminar flow, regardless of relative surface roughness, the Moody friction factor is expressed as:
f =
64
Re
Due to the inclusion of a constant temperature, electrical resistance heat sink in the
experimental setup, an estimate of the temperature drop across the epoxy was needed to complete
the model. This was accomplished by examining a derivation of Fourier’s law for onedimensional, steady state conduction:
q x = −k∇T
q x = − kA
dT kA
=
(Ts ,1 − Ts , 2 ) (W)
dx
L
with k = material thermal conductivity (W/m K), A = area (m2), L = material length (m), and Ts,1
- Ts,2 = surface temperature difference (K). For a composite wall, as in the heat sink/buried
channel chip interface, an overall heat transfer coefficient is implemented to estimate the thermal
resistance:
U=
1
Rtotal A
=
1
(W/m2 K)
Lchip
Lheat sin k
1
+ Rt ,c +
+
k heat sin k
k chip
h
with Rt,c = thermal contact resistance (0.2-0.9 m2 K/ W for 20 µm aluminum/silicon epoxy
interface [9]) and h = internal heat transfer coefficient (W/m2 K). The temperature drop across
the interface can be found from the input electric resistance heat flux, measured interface lengths,
and material properties with an expression analogous to Newton’s Law of Cooling:
q conductive = UAchip ∆T (W)
The temperature drop from the temperature controller to the internal channel surface was
estimated at 6.8-11.2 oC for the three processed channel diameters.
ii. Experimental Setup
A fluidic data acquisition system was designed and constructed to test the heat sink
capability of the buried channels. Figure 20 presents a diagram of the system, with photographs
presented in Appendix B. The system features a constant displacement gear pump, two gauges to
capture pressure drop across the chip, thermocouples inline to the flow to record fluid
temperature, two turbine flow meters to measure coolant flow, and a constant temperature bath to
regulate the inlet coolant temperature to the sample. The flow meters (25-350 mL/min sensitivity
range) were operated in series to measure the majority of the coolant flow. A portion of the
coolant through the first meter was directed to the chip through metering valves. Although the
use of an additional flow meter increases the error of the measurement, this configuration offered
the greatest flexibility of flow conditions for the initial testing phase. A polyetherimide manifold
served as an interface for the coolant flow system, aluminum heat source, and sample chip. A
17.1 W/cm2 mica flexible heater with temperature control provided thermal heat to the aluminum
heat sink, and thus the substrate of the buried channel sample. The manifold was designed to
provide a uniform fluid path to the cooling channels while blocking undesired flow around the
sample.
Figure 20: Schematic of fluidic test setup
iii. Modeled/Experimental Results
Pressure drop and thermal cooling capacity were measured for the three test samples
previously presented. We varied flow rate through each chip from 2-15 mL/min for four
substrate temperatures. These conditions were likewise entered into the iterative energy balance
model (1) to compare simulated versus experimental conditions. We performed an additional test
to determine the heat removal of the cooling fluid due to conduction from the manifold. The
experimental data was normalized based on thermal trends evident in the calibration. Figures 2123 present the experimental and simulated cooling capacity versus flow rate for Samples A (Deff =
60 µm), B (Deff = 70 µm), and C (Deff = 72 µm), respectively. Plotted data points represent an
average of 40 measured values at each condition.
The experimental data clearly fall below the simulated values cooling performance at the
higher set-point temperatures. This is most likely due to error in estimation of surface
temperature. On-chip temperature sensors would allow an accurate quantification of thermal
epoxy resistance. The experimental data does present similar trends in cooling profiles along the
surface temperature, as heat removal is directly proportional to set-point temperature. The data
also presents a convergence to a steady state cooling capacity with increasing flow rate. This
trend is expected due to the finite amount of heat conducted by the heat sink. The assumption of
fully developed conditions along the channel length also presents a source of model error. The
steady state equations were used to quantify the hydrodynamic and thermal properties in these
regions due to the unavailability of closed form undeveloped estimations.
The high flow, high surface temperature data present an additional phenomenon. Heat
flux values removed from the samples at these points are slightly greater than the theoretical
uniform maximum heat flux (24 W/cm2 for Sample A versus a theoretical 17 W/cm2). A nonuniform heat distribution was most likely because the majority of the heat sink was wrapped in
insulation.
To view the effect of cooling capacity with channel diameter, the experimental thermal
heat flux (W/cm2) at the 77 oC set-point is presented in Figure 24:
The 60 µm-effective channel diameter sample represents the most efficient heat sink of
the three samples. This is expected due to the inverse relationship of internal heat transfer
coefficient with channel diameter (Nu # = hD/k). The remaining samples, with minimal relative
difference in effective diameter, yielded similar heat removal capacity.
Figures 25-27 present experimental and simulated differential pressure drop versus flow
for the three channel geometries.
The experimental data roughly agrees with simulated results for the larger sample sizes.
Pressure drop also decreased at similar flow rates at higher temperatures with the decrease in
water density. The steep pressure curves in the small diameter sample could possibly be
explained by the estimation of the hydraulic diameter. The ratio of area to perimeter for noncircular hydraulic diameter estimation fails at geometries where the length of a profile is much
larger than the width [11]. In such cases, the hydraulic diameter is best estimated by the width of
the channel. This estimation could apply to Sample A due to the dimensions of the access trench
etch (121 µm x 33 µm, Figures 10-11). Figure 28 presents pressure versus flow for the three
experimental chips, along with the adjusted hydraulic diameter model for Sample A. The revised
estimate of hydraulic diameter clearly presents a more accurate model of differential pressure
drop.
The experimental results, along with theoretical thermal and hydrodynamic entry lengths,
are summarized:
Sample
Measured
Channel
Diameter
Effective
Diameter
Channel
Length
Modeled
Modeled
Maximum
Hydrodynamic Thermodynamic Measured
Entry Length
Entry Length
Heat
Density
(mm)
(mm)
(W/cm2)
(µm)
(µm)
(µm)
A
72.5
60.4
8.23
0.05-0.52
0.26-2.24
24.0
B
88.6
70.2
7.56
0.06-0.69
0.36-3.00
22.2
C
92.0
72.3
6.52
0.07-0.71
0.38-3.25
17.6
V. CMOS Compatibility
The final objective of this research involves the verification that the fabrication recipe
used to create buried channels does not harm the on-chip electronics performance. We
investigated the compatibility between the CMOS circuitry and the channel processing recipe by
observing the performance of a simple 120 gate clock generator circuit adjacent to a z-axis
accelerometer MEMS device. Trenches were machined 100 µm from the circuits using the
identical process steps used to create the cooling microchannels reported on above. Figure 29
displays the clock generation input and output signals prior to any fabrication of channels. A 0.5
MHz square wave output mirrors half of 0-5.25 V, 1 MHz square wave input signal. Significant
noise due to probe impedance mismatching exists in the output signal, but is not important to the
tests conducted.
Figure 30 displays the identical signal conditions for the clock generator circuit (similar
die) after the buried channel fabrication process. The clock circuit is clearly functional and
unchanged after the silicon etching and ion milling removal of the polymer.
Although future channel designs will better determine the CMOS electronics/cooling channel
compatibility, initial test results are promising and indicate that the buried channel process does
not hinder semiconductor circuit performance. This is consistent with early results from Guillou
et. al. [16], which demonstrated that released mechanical structures could be integrated 5 µm
away from CMOS circuitry with no observable change in electrical characteristics and
performance.
VI. Conclusions
This research presents a baseline fabrication process recipe for the creation of buried
microfluidic channels in bulk single-crystal silicon. We fabricated silicon heat sinks of varied
diameter using a single masking step. The samples were sealed and packaged into a fluidic data
acquisition system to test the cooling capacity of the resulting geometries. Using electric
resistance to simulate microprocessor thermal heat generation, cooling capacities up to 24.0
W/cm2 were measured for channel diameters of 72.5-92.0 µm.
The significant contribution of this effort to the goal of integrated, fluidic microprocessor
cooling lies in the use of low temperature fabrication processes to create microchannels. We
demonstrated that polymer films can be deposited as sidewall protection and serve as an etch
mask to XeF2 processing gases. While C4F8 has been used previously as sidewall protection for
the formation of scalloped undercuts in open areas [19], the derived recipe in this effort presents
the additional advantage of machining undercut channels with the use of smaller mask features
(20 µm wide trenches). Further optimization in feature reduction presents the possibility of
surface integration with CMOS electronics.
The experimental thermal analysis served to observe the effects of channel geometry on
cooling capacity. The clear goal of this research lay in the creation of buried microchannels of
10-50 µm in diameter. Smaller channel diameters produce fully developed conditions closer to
the entrance of the channel as well as increase the heat transfer from the silicon substrate to the
coolant. With greater efficiency cooling, the heat load can be satisfied with lower pumping rates,
reducing the size and power consumption of the circulating pump. The microchannel diameter
size in this analysis was directly attributed to the minimum mask feature size attainable during
processing. Significant sidewall degradation occurred when attempting to process channels
below mask sizes of 10 µm. The 20 µm mask features led to channels of 73-92 µm in diameter,
which in turn required relatively high flow rates to satisfy modest heat fluxes of ~24 W/cm2. By
optimizing the recipe to reduce feature size, higher heat capacities can be obtained at lower
pumping rates.
A number of areas of improvement have been identified for future research. They
include:
1) CMOS integration. Using either backside masking or integrating with circuitry, a
high quality aluminum mask would eliminate photolithography error and feature
expansion. On chip polysilicon resistance heaters would eliminate the need for
thermal bonding as well as offer high thermal heat generation. CMOS temperature
sensors could give an accurate substrate temperature profile during cooling.
2) Feature reduction. In addition to integration and cooling efficiency, smaller feature
sizes offer greater flexibility in sealing the access trenches. This process was proven
successfully with 20 µm trenches, but further optimization could determine the
feasibility of 5 µm trenches as well as access holes.
3) Additional low temperature passivation. The STS ICP polymer deposition rate was
measured to vary up to 15% for extended 8.5 minute processes. Due to the porous
nature of the passivation, XeF2 gas was also able to penetrate through the silicon in
thin regions of the sidewalls. Future research could determine the worthiness of
lower temperature growth (~400 oC) SiO2 as a sidewall mask.
4) Boiling. Phase change heat transfer will be needed to successfully cool
microprocessors with thermal heat loads up to 300 W/cm2. This leads to the need for
microscopic inspection of bubbles at the outlet header of the test chip, as well as
improved packaging.
5) Customized flow meter. The flow divider system used in the experimental setup was
selected in order to accommodate the high flow requirements of the flow meters.
Custom meters with sensitivity ranges as low as 30 mL/min would offer greater
accuracy and simplicity.
6) Custom packaging: The manifold used to house the test chip was driven in part by the
need to house the electric resistance heater. Packaging with CMOS test chips would
eliminate heat transfer to the fluid through the manifold and allow wire bonding at
the chip surface.
VII. References
[1]
Tuckerman, D. B. and Pease, R. F. W., “High-Performance Heat Sinking for VLSI,”
IEEE Electron Device Letters, vol. EDL-2, no. 5, pp.126-129, 1981.
[2]
Bergles, A. E., Kendall, G. E., Griffith, P., and Lienhard V, J. H., “Small Diameter
Effects on Internal Flow Boiling,” Proceedings of ASME 2001 IMECE, pp. 1-17, 2001.
[3]
de Boer, M. J., Tjerkstra, R. W., Berenschot, J. W., Jansen, H. V., Burger, G. J.,
Gardeniers, J. G. E., Elwenspoeck, M., van den Berg, A., “Micromachining of Buried Micro
Channels in Silicon,” Journal of Microelectromechanical Systems, vol. 9, no. 1, pp. 94-103, 2000.
[4]
Joo, Y., Dieu, K., Kim, C., “Fabrication of Monolithic Microchannels for IC Chip
Cooling,” Micro Electro Mechanical Systems, 1995, MEMS ’95, Proceedings IEEE, pp. 362-367.
[5]
Vieder, C., Holm, J., Forssén, L., Elderstig, H., Lindgren, S., Åhlfeldt, H., “A New
Process for Combining Anisotropic Bulk Etching with Subsequent Precision
Lithography,”Transducers ’97, 1997 International Conference on Solid-State Sensors and
Actuators, pp. 679-682.
[6]
Perret, C., Boussey, J., Schaeffer, Ch., Coyount, M., “Integration of Cooling Devices in
Silicon Technology,” 5th International Workshop Thermal Investigations of ICs and Systems, pp.
1780-1786, 1999.
[7]
Zhimin, W., Fah, C., “The Optimum Thermal Design of Microchannel Heat Sinks,” 1997
IEEE/CPMT Electronic Packaging Technology Conference, pp. 123-129, 1997.
[8]
Jiang, X. N., Zhou, Z. Y., Huang, X. Y., Liu, C. Y., “Laminar Flow Through
Microchannels Used for Microscale Cooling Systems,” 1997 IEEE/CPMT Electronic Packaging
Technology Conference, pp. 119-122, 1997.
[9]
Incropera, F. P., De Witt, D. P., “Fundamentals of Heat and Mass Transfer,” pp. 79-92,
312-335, 467-494, 3rd Edition, John Wiley and Sons, New York, 1990.
[10]
Shames, I. H., “Mechanics of Fluids,” pp. 709-712, 3rd Edition, McGraw-Hill Inc., New
York, 1992.
[11]
Crane, “Flow of Fluids Through Valves, Fittings, and Pipe, Technical Paper No. 410,”
pp. 1.1-1.7, 21st Edition, Crane Co. 1982.
[12]
Azar, K, and Morabito, J., “Managing Power Requirements in the Electronics
Industry,”Electronics Cooling, Vol. 6, No. 4, December 2000.
[13]
Lucent Technology Microelectronics Databook
[14]
Allen, A., “Digital Silicon Technology,” NEMI Workshop, VA.
[15]
Datasheet: Intel Pentium 4 Processor in the 478-Pin Package: 1.5-2 GHz.
[16]
Guillou, D. E., Santhanam, S., and Carley, L. R., “Laminated, Sacrificial-Poly MEMS
Technology in Standard CMOS,” Sens. Actuators A, Phys(Switzerland), Sensors and Actuators A
(Physical), vol. A85, no. 1-3, pp. 346-55, 1999.
[17]
Recommended by S. Santhanam (Carnegie Mellon University), 11/13/2001.
[18]
Recommended by J. Zahn (Carnegie Mellon University MEMS Laboratory), 7/1/2001.
[19]
First demonstrated by L. Erdmann (Carnegie Mellon University MEMS Laboratory) and
C. Wu (Carnegie Mellon University M.E. Dept.) in the creation of bulk-silicon scallops for nozzle
turbulence enhancement, 10/6/2000.
[20]
Recommended by STS support personnel, 8/2001.
[21]
Maluf, N., “An Introduction to Microelectromechanical Systems Engineering,” pp. 6670, Artech House, Inc., 2000.
VIII. Appendix A
Microchannel Processing Recipe
4 um SiO2 silicon wafer, ~1.5 cm x 1.5 cm cleaved samples
Step
Time (min)
Pressure
(mT)
Flow (sccm)
1-2
Photolithography
(Karl Suess
MA56)
3
SiO2 etch (PTI)
112
125
4
DRIE (STS)
Etch: 22;
Pass: 12
5
Isotropic (STS)
70: 12/8 sec
etch/passivate
cycle
3
6
Passivation
(STS)
Ion milling
(STS)
RIE (STS)
8.5
12
CHF3: 22.5;
N2: 5
Etch: SF6: 130,
O2: 13;
Pass.: C4F8: 85
SF6 :130,
C4F8: 18
C4F8: 85
2-3
1.4
30
0.3
22
SF6: 130, O2:
13
9a
9b
Oven cure
Isotropic (Xactix
XeF2)
10
Oxygen plasma
(PTI)
Wafer bonding
30
A: 20
B: 23
C: 33
20
7
8
11
Power (W)
Additional
AZ 4620, 3K rpm, 120 sec.
110 oC hot plate, 17 sec. exp
@ 14.1 mJ/cm2, 60 min.
120 oC oven hard bake
32
100
Etch: 600 coil,
12 platen;
Pass.: 600 coil
600 W coil, 3
W platen
600 W coil
700 W coil,
150 W platen
600 coil, 12
platen
Etch rate: 2-2.5 µm/minute
Deposition rate: 0.1-0.12
µm/min
Platen bias voltage: 80-130
V
T = 200 oC
60 sec. cycles
Xe: 2
N2: 10
150
Etch rate: 2-2.5 µm/minute
50
100
Epotek 302-3M, 4K rpm, 2
hr. 60 oC cure
Download